Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

The Astrophysical Journal, 732:82 (11pp), 2011 May 10 doi:10.

1088/0004-637X/732/2/82

C 2011. The American Astronomical Society. All rights reserved. Printed in the U.S.A.

CONSTRAINING THE DARK ENERGY EQUATION OF STATE USING LISA OBSERVATIONS OF SPINNING
MASSIVE BLACK HOLE BINARIES
Antoine Petiteau, Stanislav Babak, and Alberto Sesana
Max-Planck-Institut fuer Gravitationsphysik, Albert-Einstein-Institut, Am Muhlenberg 1, D-14476 Golm bei Potsdam, Germany
Received 2010 December 10; accepted 2011 March 1; published 2011 April 20

ABSTRACT
Gravitational wave (GW) signals from coalescing massive black hole (MBH) binaries could be used as standard
sirens to measure cosmological parameters. The future space-based GW observatory Laser Interferometer Space
Antenna (LISA) will detect up to a hundred of those events, providing very accurate measurements of their luminosity
distances. To constrain the cosmological parameters, we also need to measure the redshift of the galaxy (or cluster of
galaxies) hosting the merger. This requires the identification of a distinctive electromagnetic event associated with
the binary coalescence. However, putative electromagnetic signatures may be too weak to be observed. Instead, we
study here the possibility of constraining the cosmological parameters by enforcing statistical consistency between
all the possible hosts detected within the measurement error box of a few dozen of low-redshift (z < 3) events.
We construct MBH populations using merger tree realizations of the dark matter hierarchy in a ΛCDM universe,
and we use data from the Millennium simulation to model the galaxy distribution in the LISA error box. We show
that, assuming that all the other cosmological parameters are known, the parameter w describing the dark energy
equation of state can be constrained to a 4%–8% level (2σ error), competitive with current uncertainties obtained
by type Ia supernovae measurements, providing an independent test of our cosmological model.
Key words: black hole physics – cosmological parameters – galaxies: distances and redshifts – gravitational waves
– methods: statistical
Online-only material: color figures

1. INTRODUCTION in order to measure the source redshift and be able to do


cosmography. If the event is nearby (z < 0.4), then we have
The Laser Interferometer Space Antenna (LISA; Danzmann & a very good localization of the source on the sky and we can
the LISA Study Team 1997) is a space-based gravitational wave identify a single cluster of galaxies hosting the merger. As we
(GW) observatory which is expected to be launched in 2022+. go to higher redshifts, LISA sky localization abilities become
One of its central scientific goals is to provide information quite poor: a typical sky resolution for an equal mass 106 M
about the cosmic evolution of massive black holes (MBHs). It inspiralling MBH binary at z = 1 is 20–50 arcmin per side at
is, in fact, now widely recognized that MBHs are fundamental 2σ (Trias & Sintes 2008; Lang & Hughes 2009; Arun et al.
building blocks in the process of galaxy formation and evolution; 2009b), which is in general not sufficient to uniquely identify
they are ubiquitous in nearby galaxy nuclei (see, e.g., Magorrian the host of the GW event. There is, therefore, a growing interest
et al. 1998), and their masses tightly correlate with the properties in identifying putative electromagnetic signatures associated
of their host (Gültekin et al. 2009, and references therein). In with the MBH binary before and/or after the final GW-driven
popular ΛCDM cosmologies, structure formation proceeds in a coalescence (for a review, see Schnittman 2010, and references
hierarchical fashion (White & Rees 1978), through a sequence therein). Electromagnetic anomalies observed before or after
of merging events. If MBHs are common in galaxy centers at the coalescence within the LISA measurements error box may
all epochs, as implied by the notion that galaxies harbor active allow us to identify the host and to make a redshift measurement.
nuclei for a short period of their lifetime (Haehnelt & Rees However, most of the proposed electromagnetic counterparts
1993), then a large number of MBH binaries are expected to are rather weak (below the Eddington limit), and in case of
form during cosmic history. LISA is expected to observe the dry mergers (no cold gas efficiently funneled into the remnant
GW-driven inspiral and final coalescence of such MBH binaries nucleus) we do not expect any distinctive electromagnetic
out to very high redshift with a high signal-to-noise ratio (S/N), transient. This brings us back to the original idea by Schutz
allowing very accurate measurements of the binary parameters. (1986) of considering each galaxy within the LISA measurement
The collective properties of the set of the observed coalescing error box as a potential host candidate. The idea is that, by
binaries will carry invaluable information for astrophysics, cross-correlating several GW events, only one galaxy (cluster
making it possible to constrain models of MBH formation and of galaxies) in each error box will give us a consistent set of
growth (Plowman et al. 2010; Gair et al. 2010; Sesana et al. parameters describing the universe. The effectiveness of this
2010). method has been demonstrated by MacLeod & Hogan (2008) in
Besides astrophysical applications, coalescing MBHs could the context of the Hubble constant determination by means of
be used as standard sirens (Schutz 1986; Holz & Hughes 2005; low-redshift (z < 0.2) extreme mass ratio inspirals.
Lang & Hughes 2006, 2009; Arun et al. 2007, 2009a; Van Den We use the hierarchical MBH formation model suggested by
Broeck et al. 2010). The high strength of the GW signals allows Volonteri & Begelman (2010) to generate catalogs of coalescing
us to measure the luminosity distance with a precision of less MBH binaries along the cosmic history. This model predicts
than a percent at redshift z = 1 (neglecting weak lensing). ∼100 MBHs mergers observable by LISA in three years, in
However, we need an electromagnetic identification of the host the redshift range [0:5]. We do not use sources beyond redshift

1
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

z = 3 due to difficulties of measuring galaxy redshifts beyond 2. ANALYTICAL FRAMEWORK


that threshold.1 We model the galaxy distribution in the universe
using the Millennium simulation (Springel et al. 2005). For each 2.1. Cosmological Description of the Universe
coalescing MBH in our catalog, we select a host galaxy in the We assume the standard ΛCDM model, which describes
Millennium run snapshot closest in redshift to the actual redshift our universe as the sum of two non-interacting components:
of the event. For each galaxy in the snapshot, we compute (1) a pressureless component corresponding to all visible and
the apparent magnitude in some observable band, and we DM, and (2) a dark energy component with current effective
create a catalog of redshift measurements of all the observable equation of state ω(z) (for the standard Λ-term, p = −,
potential host candidates. Note that typical observed mergers it would correspond to ω = −1). Current estimates based
involve 104 –106 M MBHs, which implies (using the black hole on SN Ia observations and anisotropy measurements in the
mass–bulge relations; see, e.g., Gültekin et al. 2009) relatively cosmic microwave background (Riess et al. 1998; Komatsu
light galaxies. However, observed galaxies are heavy due to et al. 2011) tell us that about 70% of the universe en-
selection effects: roughly speaking, mass reflects luminosity, ergy content is in the form of the dark energy. The evolu-
so that at high redshifts we can observe only very massive tion of the universe is therefore described by the expansion
(luminous) galaxies. Therefore, the actual host might not be (and equation
often is not) among the observed galaxies. The important fact is
the self-similarity of the density distribution: the local density    z 
1 + ω(z)
distribution for all galaxies and the density distribution for heavy H = H0 Ωm (1 + z) + Ωde exp 3
2 2 0 3 0
dz , (1)
galaxies are quite similar, which allow us to infer the likelihood 0 1+z
of the host redshift on the basis of redshift measurements of the
where H = ȧ/a (a being the lengthscale of the universe) is the
luminous galaxies only.
Hubble expansion parameter and H0 is its current value (t = 0),
We assume that the GW source parameter measurements (GW
Ωm and Ωde are the ratios of the matter density and the dark
likelihoods) are represented by multivariate Gaussian distribu-
energy density to the critical density, and ω(z) describes the
tions around the true values, with the variance–covariance ma-
effective dark energy equation of state as a function of z. We
trix defined by the inverse of the Fisher matrix. This is a good
assume that the universe is spatially flat, the luminosity distance
approximation in the case of Gaussian instrumental noise and
is therefore computed as
large S/N. At z  0.25, the uncertainty in the luminosity dis-
tance (DL ) is dominated by weak lensing (WL) due to the ex-  z
dz
tended distribution of dark matter (DM) halos between us and DL = (1 + z) . (2)
the GW source. In this paper, we combine the luminosity dis- 0 H (z )
tance errors given by GW measurements and WL, referring to
them as GW+WL errors. We use two estimations of the WL In our simulations, we fix all parameters (assuming that they
error (1) from Shapiro et al. (2010) and (2) from Wang et al. are known exactly) to the currently estimated mean values:
(2002). H0 = 73.0 km s−1 Mpc−1 , Ωm = 0.25, and Ωde = 0.75.
In order to evaluate the error box, we need to assume some We also simplify the form of ω(z) for which we will assume
prior on the cosmological parameters. In this exploratory study, ω = −1 − w, where w is a constant.2 We choose the value
we assume that we know all the cosmological parameters but w = 0 to simulate our universe which is what has been used in
the effective equation of state for the dark energy, described the Millennium simulation (see below).
by the parameter w (which could be the case by the time LISA
will fly). In a follow-up paper, we will relax this assumption 2.2. Methodology and Working Plan
by including also the Hubble constant and the matter and dark Our aim is to show that we can constrain w via GW obser-
energy content of the universe as free parameters. We take the vations of spinning MBH binaries using a Bayesian framework.
prior range for w from the seven year Wilkinson Microwave Let us consider j = 1, . . . , Nev GW observations. For each
Anisotropy Probe (WMAP) analysis (Komatsu et al. 2011). We event, we can infer the probability of a parameter w, given the
show that using statistical methods w can be constrained to collected data s, using Bayes theorem:
a 4%–8% level (2σ error), providing an effective method for
estimating the dark energy equation of state. We also show that p0 (w)Pj (s|w)
this result depends weakly on the prior range and could serve as Pj (w|s) = . (3)
Ej
an independent way of measuring the dark energy equation of
state, with respect to canonical methods employing observations
Here, Pj (w|s) is the posterior probability of the parameter
of type Ia supernovae (SNe Ia; Riess et al. 1998).
w, Pj (s|w) is the likelihood of the parameter w given the
The paper is structured as follows. In Section 2, we explicitly
observation s, p0 (w) in the prior knowledge of w and Ej is
spell out all the details of the adopted cosmological model
defined as 
and of the Bayesian analytical framework. In Section 3, we
give more insights on the MBH population model and on the Ej = p0 (w)Pj (s|w)dw. (4)
galaxy distributions extracted from the Millennium database. In
Section 4, we describe our simulated GW and electromagnetic The likelihood Pj (w|s) must be appropriately specialized to
observations. We give results of our simulations under different our problem. We want to exploit GW observations to constrain
assumptions about WL, depth of the follow-up electromagnetic w through the distance–redshift (DL –z) relation as given by
surveys, etc., in Section 5. We summarize our findings in Equation (2).
Section 6.
1 There are other reasons for not going beyond z = 3 which we will discuss 2 Here, we use notations for the dark energy equation of state adopted in the
later. WMAP data analysis (Komatsu et al. 2011).

2
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

1. The distance DL is provided by the GW observations: the all Nev GW events are independent, the combined posterior
GW signal carries information about the parameters of the probability is
binary, including its location on the sky and its luminosity  ev
distance. All of those parameters can be extracted using p0 (w) Nj =1 Pj (s|w)
latest data analysis methods (Petiteau et al. 2010; Cornish P (w) =  Nev . (7)
p0 (w) j =1 Pj (s|w)dw
& Porter 2007). The measurements errors are encoded in
the GW likelihood3 function L(DL , θ, φ, λ), where {θ, φ} To evaluate w through Equation (7) we therefore need the
are the ecliptic coordinates of the source and λ  represents following.
all of the other parameters characterizing the MBH binary 1. An MBH binary population model defining the properties
(spin and orientation, mass, orientation of the orbit, and of the Nev coalescing systems;
the MBH’s position at the beginning of observations). 2. The spatial distribution of galaxies within a volume com-
When estimating DL , WL cannot be neglected. In fact, parable with the combined GW+WL measurement error
the error coming from the WL (causing fluctuations in the box;
brightness of the GW source, which gives an uncertainty 3. The measurement errors associated with GW observations
in the luminosity distance) dominates over the GW error 
of coalescing MBH binaries (defining Lj (DL , θ, φ, λ));
starting from redshift z ∼ 0.25 (see Figure 2). 4. An estimation of spectroscopic survey capabilities to con-
2. The redshift measurement does not rely on any distinctive struct the galaxy redshift distribution within the GW+WL
electromagnetic signature related to the GW event. We measurement error box (defining pj (θ, φ, z)).
extract a redshift probability distribution of the host from
the clustering properties of the galaxies falling within the We will consider these points individually in the next two
GW+WL error box. This defines an astrophysical prior sections.
p(θ, φ, z) for a given galaxy in the measurement error
3. ASTROPHYSICAL BACKGROUND
box to be the host of coalescing binary. To translate the
measured DL and uncertainty ΔDL of the GW event into 3.1. Massive Black Hole Binary Population
a corresponding z and Δz for the candidate host galaxies
in the sky, we use the prior knowledge of p0 (w) obtained To generate populations of MBH binaries in the universe,
from WMAP. we use the results of merger tree simulations described in de-
tails in Volonteri et al. (2003). MBHs grow hierarchically, start-
The likelihood in Equation (3) can therefore be written as ing from a distribution of seed black holes at high redshift,
 through a sequence of merger and accretion episodes. Two dis-
 ]p(λ
 )pj (θ, φ, z) tinctive type of seeds have been proposed in the literature. Light
Pj (s|w) = Lj [DL (z, w), θ, φ, λ
(M ∼ 100 M ) seeds are thought to be the remnant of Popu-
 dθ dφ dz, lation III (POPIII) stars (Madau & Rees 2001), whereas heavy
× dλ (5)
seeds form following instabilities occurring in massive proto-
 on the parameters λ  galactic disks. In the model proposed by Begelman Volonteri
where we have introduced the priors p(λ) & Rees (Begelman et al. 2006, hereafter BVR model), a
(which we assume in this paper to be uniform). It is convenient “quasistar” forms at the center of the protogalaxy, eventually
to change the variable of integration from DL to z. Since we have collapsing into a seed BH that efficiently accretes from the qua-
assumed uniform priors on λ,  we can marginalize the likelihood sistar envelope, resulting in a final mass M ∼ few ×104 M .
over those parameters4 to obtain Here, we use the model recently suggested by Volonteri &
 Begelman (Volonteri & Begelman 2010, hereafter VB model),
Pj (s|w) = πj [DL (z, w), θ, φ] pj (θ, φ, z) dθ dφ dz, (6) which combines the two above prescriptions by mixing light
and heavy initial seeds. This model predicts ∼30–50 events per
year in the redshift range 0 < z < 3, relevant to this study.
where we denoted the marginalized GW likelihood as The dashed blue lines in Figure 1 show the redshifted total mass
πj [DL (z, w), θ, φ]. Practically, we limit the integration to the (Mz = (M1 + M2 )(1 + z), being M1 > M2 the rest-frame masses
size of the error box (in principle the integration should be taken of the two MBHs, upper left panel), mass ratio (q = M2 /M1 ,
over the whole range of parameters but we found that consider- upper right panel), and redshift (lower panel) distribution of
ing the 2σ error box is sufficient). the MBH binaries coalescing in three years, as seen from the
We assume that the error in luminosity distance from the WL Earth. The model predicts ∼40 coalescences in the redshifted
is not correlated with the GW measurements, hence the integral mass range 105 M < Mz < 107 M , almost uniformly dis-
in Equation (6) can be performed over the sky ({θ, φ}) first, and tributed in the mass ratio range 0.1 < q < 1, with a long
then over the redshift. We also found that the correlation between tail extending to q < 10−3 . For comparison, we also show the
DL and the sky position coming from the GW observations is population expected by a model featuring a heavy seed only
not important for events at z < 0.5. Plugging Equation (6) (BVR model, green dot-dashed lines), and by an alternative
into Equation (3) defines the posterior distribution of w for a VB type model (labeled VB-opt for optimistic, red long-dashed
single GW event (as indicated by the index j). Assuming that lines) with a boosted efficiency of heavy seed formation (see
Volonteri & Begelman 2010 for details). It is worth mention-
3 ing that these models successfully reproduce several properties
Through the paper, with GW likelihood we mean the likelihood of the LISA
data to contain the GW signal with a given parameters, not to be confused with of the observed universe, such as the present-day mass density
the likelihood Pj (s|w) defined in the Bayes theorem. of nuclear MBHs and the optical and X-ray luminosity func-
4 Here, this corresponds to the projection of the Fisher matrix to
tions of quasars (Malbon et al. 2007; Salvaterra et al. 2007).
three-dimensional parameter space of sky location θ, φ, and luminosity The BVR and the VB-opt models predict MBH population ob-
distance DL .
servables bracketing the current range of allowed values. The
3
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

the GW signal according to a probability proportional to the


number density of neighbor galaxies ngal . Such an assumption
comes from the fact that two galaxies are needed in order to
form an MBH binary, and we consider that the probability that a
certain galaxy was involved in a galaxy merger is proportional to
the number of neighbor galaxies. We consider to be neighbors of
a specific galaxy all the N (R) galaxies falling within a distance

R = σ TH (z), (8)

where σ = 500 km s−1 is the typical velocity dispersion


of galaxies with respect to the expanding Hubble flow, and
TH (z) is the Hubble time at the event redshift. The number
density of neighbor galaxies is then simply written as ngal =
3N (R)/(4π R 3 ). When we choose the merger host, we compute
ngal considering all the neighbor galaxies, without imposing any
kind of mass or luminosity selection. In this case ngal ≡ ntotal .
However, when we will construct the probability of a given
observable galaxy to be the host of the merger (i.e., the
Figure 1. Population of coalescing MBH binaries in three years. Top left panel: astrophysical prior pj (θ, φ, z)), we will have to compute ngal
total redshifted mass distribution; top right panel: mass ratio distribution; lower according to the number of observed neighbors, because this
panel: redshift distribution. Color and linestyle codes are labeled in the figure. is the only thing we can do in practice when we deal with an
(A color version of this figure is available in the online journal.) observed sample of galaxies (see Section 4.2).

4. SIMULATING THE OBSERVATIONS


VB-opt model, in particular, is borderline with current observa-
tional constraints on the unresolved X-ray background, and it is 4.1. Gravitational Wave Observations: Shaping the Error Box
shown here only for comparison. In the following, we consid-
ered the VB model only, which fits all the relevant observables As we mentioned in Section 3.1, we performed 100 realiza-
by standing on the conservative side. tions of the MBH binary population from the VB model. Each
We performed 100 Monte Carlo realizations of the population realization contains 30–50 events in the redshift range [0:3]. The
of MBH binaries coalescing in three years. Each realization total mass, mass ratio, and redshift distributions of the events
takes into account the distribution of the number of events and are shown in Figure 1. In order to simulate GW observations,
MBH masses with the redshift as predicted by the VB model. the binary sky location is randomly chosen according to a uni-
Other parameters (like time of coalescence, spins, initial orbital form distribution on the celestial sphere, the coalescence time
configuration) are chosen randomly using uniform priors over is chosen randomly within the three years of LISA operation
the appropriate allowed ranges. (we assume three years as default mission lifetime). The spin
magnitudes are uniformly chosen in the interval [0:1] in units
of mass squared, and the initial orientations of the spins and of
3.2. Galaxy Distribution the orbital angular momentum are chosen to be uniform on the
To simulate the galaxy distribution in the universe, we use sphere. More detailed description of the model for GW signal
the data produced by the Virgo Consortium publicly available used in this paper is given in Petiteau et al. (2010).
at http://www.g-vo.org/Millennium. These data are the result of The GW likelihood L needed in Equation (5) is approximated
the implementation of semianalytic models for galaxy formation as a multivariate Gaussian distribution with inverse correlation
and evolution into the DM halo merger hierarchy generated matrix given by the Fisher information matrix (FIM):
by the Millennium simulation (Springel et al. 2005). The
L ∼ e−(s−h|s−h) ∼ e(θ −θ̂ i )Γij (θ j −θ̂ j )/2
i
Millennium run is an N-body simulation of the growth of DM . (9)
structures in the expanding universe starting from a Gaussian
spectrum of initial perturbations in the DM field at high redshift, Here, θ i is the vector of the parameters characterizing the
which successfully reproduced the net-like structure currently spinning MBH binary, θ̂ i are the maximum likelihood estimators
observed in the local universe. The simulation has a side for those parameters which are assumed to correspond to
length of ≈700 Mpc (comoving distance), and its outcome is the true values (no bias), and Γij = (h,i |h,j ) is the FIM,
stored in 63 snapshots evenly separated in log(z), enclosing where the commas correspond to derivatives with respect to
all the properties of the DM structure at that particular time. the parameters. This is a reasonable approximation due to the
Semianalytical models for galaxy formation are implemented a large S/N (for more details on the FIM and its applicability,
posteriori within the DM structures predicted by the simulation. see Vallisneri 2008). Our uncertainties on estimated parameters
Such models have been successful in reproducing several are consistent with Lang & Hughes (2009), Babak et al. (2010),
observed properties of the local and the high-redshift universe and Petiteau et al. (2010). We did not include higher harmonics
(see, e.g., Bower et al. 2006; De Lucia & Blaizot 2007). Here, we (only the dominant, twice the orbital frequency) as they only
use the implementation performed by Bertone and collaborators slightly improve parameter estimation for precessing binaries.
(Bertone et al. 2007), which is a refinement of the original However, including higher harmonics in the GW signal model is
implementation by De Lucia & Blaizot (2007). important in the case of the small spins and low precession (when
For each coalescing MBH binary, we choose the snapshot spins are almost (anti)aligned with the orbital momentum; Lang
closest in redshift. Within the snapshot we choose the host of et al. 2011). We use truncated waveforms corresponding to the

4
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

0.06

0.05
GW error
Weak lensing error #1
Weak lensing error #2
Combined error #1
0.04 Combined error #2
δDL/DL

0.03
400
350
300
0.02 250
400
200
150 350
100
50 300
0.01
250

50 200
0 100
0 0.5 1 1.5 2 2.5 3
z 150 150
200
Figure 2. Relative error in the luminosity distance due to WL from 250 100
300
(1) Shapiro et al. (2010, circles) and from (2) Wang et al. (2002, squares). The 350
black solid line is the median error due to GW measurements only; the solid- 400 50
circle and the solid-square lines are for the combined errors under assumptions
Figure 3. Example of error box (cylinder) in part of the Millennium snapshot
(1) and (2), respectively (see the text).
(cube with unit in Mpc). The blue cylinder is the measurement error box and
(A color version of this figure is available in the online journal.) the green one also considers the prior on w. The black big dot is the host and
the brown small dots are the selected galaxy candidates.
(A color version of this figure is available in the online journal.)
inspiral only. However, the addition of a merger and ringdown
will further reduce the localization error due to the higher S/N 2. We pick a galaxy (red dot) in the snapshot with a probability
(McWilliams et al. 2010). This error is usually an ellipse on given by the local galaxy number density ntotal .
the sky but we simplify it by choosing the circle with the same 3. We construct around the galaxy an error box given by ΔΩ
area. and ΔzGW+WL , and the galaxy can lie anywhere with respect
For the luminosity distance measurement, we need to take into to this error box (blue cylinder).
account the WL. We assume the WL error to be Gaussian with a 4. We expand the error box along the direction of the observer
σ given by (1) Shapiro et al. (2010). Such assumption is rather both sides by Δz given by the uncertainty in w (green
pessimistic; we also tried the prescription given by (2) Wang cylinder).
et al. (2002), which gives smaller errors, but still larger than the 5. According to some prescription, which we will describe in
level that may be achieved after mitigation through shear and the next section, we select observable galaxies in the error
flexion maps (Hilbert et al. 2011). Both of those estimations are box (brown dots).
represented in Figure 2 as (1) dark (red online) circles and (2)
light (orange online) squares correspondingly. The median error As shown in Figure 3, we interpret one of the directions
in DL due to GW measurements only is given by the solid black in the Millennium snapshot as distance from the observer, and
line. The combined error for model (1) is given by the upper convert the comoving distance in redshift. We assume a periodic
(blue) circle-line curve, and for model (2) by the lower (green) expansion of the Millennium data in order to fit large error boxes.
square-line curve. We consider our setup to be conservative in Note that the original Millennium simulation also assumes the
the estimation of the WL effects. The main aim of this work same periodicity in the distribution of the matter. The size of
is to build a reasonable setup for what could be observed by the error box at high redshift covers a significant fraction of
the time LISA will fly, and make a first order estimation of the simulation box so we do not go beyond the redshift z = 3
LISA capabilities to constrain the dark energy equation of state. (as we will show later, spectroscopic observations at such high
We will address non-Gaussianity of the WL as well as other redshifts will be impractical anyway). Together with larger error
corrections to the model to make it more realistic in a follow-up boxes, we have a nonlinear increase in the number of events
paper. at high redshift. To reduce the overlap between error boxes
We consider an error box size corresponding to 2σ of the corresponding to different GW events, we choose cylinders with
measurement errors in the sky location (σsky ) and in the source random orientations.
distance as evaluated by the FIM plus WL uncertainties. For Figure 4 shows an example of the resulting weighted distri-
observational purposes, the dimensions of this error box are bution of galaxy redshifts (with weight proportional to the local
ΔΩ = 2σsky and Δz. For the latter, we also include the density ntotal ). It is a projection of the clumpiness along the line
uncertainty given in the Dl –z conversion due to the error (prior) of sight which is also proportional to the probability distribution
on w, p0 (w). of z for the event. The probability distribution of w for the event
Let us summarize how we construct an error box in practice, will be directly related to this result. We noticed that there is a
as, for example, the one illustrated in Figure 3. very large number of underdense regions and several very dense
superclusters. The probability of a galaxy with a low local den-
sity to host a merger is very low but there is a huge number of
1. We select the closest Millennium snapshot to the event in such galaxies, and we found that the probability of the host to
redshift. be in (super)clusters is similar to that of being in a low-density

5
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

3 tering information we need by getting spectra of the brightest


objects only.
At z = 1, the typical number of galaxies enclosed in the 2σ
2.5
error box described above is in the range 104 –105 . However,
not all of them are bright enough to get useful spectra. The
2 semianalytic galaxy evolution model (Bertone et al. 2007)
scaled local density

implemented on top of the Millennium run returns the stellar


mass of each galaxy, and the absolute bolometric magnitude
1.5 Mb . By knowing the redshift, and by using standard galactic
templates one can therefore compute the apparent magnitude in
a given band, by assuming the appropriate k-correction (Oke &
1
Sandage 1968). Here, we use the R-band apparent magnitude
mr for illustrative purposes, and we adopt the relation (Zombeck
0.5 1990)
Mb = −5log(zc/H0 ) − 1.086z − 25 + mr + 0.6, (10)
0
1.2 1.25 1.3 1.35 1.4 1.45 1.5 where 0.6 is the k-correction. For each galaxy, we compute mr
z
and we simulate spectroscopic surveys at different thresholds
Figure 4. Distribution of the weighted galaxies with the redshift. The green
dashed vertical line is the redshift of the host galaxy.
mr = 24, 25, 26. We stress here that the GW host was
chosen among all of the galaxies falling in the error box, and
(A color version of this figure is available in the online journal.)
therefore may not (and usually does not) belong to the observed
sample. We then assume that for each galaxy satisfying the
region. As we will see later in the result section, this may cause survey threshold we get an exact spectroscopic redshift, and we
a very wrong estimation of w for some individual GW event. combine the redshift distribution of several error boxes to get a
statistical estimation of w. In practice, each redshift estimation
will come with a measurement error, and an intrinsic error due
4.2. Redshift Measurements Through Spectroscopic Surveys to the proper motion of the source with respect to the Hubble
flow. Both errors are, however, of the order of Δz/z < 10−3 ,
To get a statistical measurement of w, we need to exploit the well below the typical redshift scale corresponding to spatial
clustering of the galaxies falling within the error box (which clustering of structures (Δz ∼ 0.01; see Figure 4) we need to
defines the astrophysical prior pj (θ, φ, z) in Equation (5)). It resolve.
is therefore necessary to get efficient redshift measurements of Our method does not rely on the observation of a prompt
thousands of galaxies within a small field of view (FOV): the in- transient associated with the MBH binary coalescence to iden-
formation we seek is enclosed in the redshift distribution of such tify the host galaxy. Nevertheless, getting thousands (or tens
galaxies. We stress here that we are not looking for a distinctive of thousands) of spectra in a small FOV requires a dedicated
electromagnetic counterpart to the GW event. In fact, the actual observational program. Thanks to multi-slit spectrographs such
host of the coalescing binary may not even be observable. Typi- as VIMOS at Very Large Telescope (VLT; Le Fèvre et al. 2003)
cal masses of our binaries are 105 –106 M . Using MBH–bulge and DEIMOS at Keck (Faber et al. 2003), fast deep spectro-
scaling relations (Gültekin et al. 2009), such MBHs are expected scopic surveys of relatively large FOV are now possible. For
to be hosted in galaxies with stellar mass 109 –1010 M , i.e., in example, the ongoing VIMOS VLT deep survey (Le Fèvre et al.
DM halos with total mass <1011 M . The Millennium run mass 2005), took spectra of >10,000 galaxies, mostly in the redshift
resolution is ∼109 M , meaning that typical host structures are range 0 < z < 1.5, within an FOV of 0.61 deg2 at an apparent
formed by less than 100 particles. Unfortunately, the Millen- magnitude limit IAB < 24. Comparable figures are achieved by
nium run is severely incomplete in the expected mass range of other observational campaigns such as zCOSMOS (Lilly et al.
LISA MBH binary hosts. Here, we do not attempt to exploit any 2009) and DEEP2 (Davis et al. 2003), which were able to survey
MBH–host relation to select the host of our GW event; the prob- selected galaxies in various photometric bands (U, B, R, I ) to
ability of being a host is only related to the local number density an apparent magnitude limit of about 24. Going deeper in red-
of neighbor galaxies ntotal . Such an assumption relies on the con- shift, Lyman break galaxy redshift surveys are finding success in
cept of self-similarity of the galaxy clustering at different mass efficiently getting high-quality spectra of hundreds of galaxies
scales: typical LISA MBH binary hosts cluster in the same way as in the redshift range 2.5 < z < 3.5 within an FOV ∼ 1 deg2
more massive galaxies. We checked this assumption by compar- (Bielby et al. 2010). To illustrate this, the VIMOS spectrograph
ing the spatial distribution of galaxies in different mass ranges can take ∼500 high-quality spectra per pointing with an inte-
(109 –1010 M , 1010 –1011 M , 1011 –1012 M ), within simula- gration time of about 4 hr, within a 7 × 8 arcmin2 FOV, which
tion snapshots at different redshift, and we postulate that this is coincidentally of the same order of the typical error box for
self-similarity extends to lower masses, below the Millennium a z = 1 GW event. The typical redshift accuracy of the spectra
run resolution. This point is crucial for two reasons: (1) espe- is Δz < 10−3 (3 × 10−4 in the zCOSMOS survey, 2 × 10−3 in
cially at z > 1, we will be able to get only spectra of luminous the Lyman break galaxy survey), well below the typical redshift
(massive) galaxies, and we need to be confident that their spatial scale we are interested in (z ∼ 0.01).
distribution mimics that of lighter galaxies that may host the GW Such results are a testament to the feasibility of efficient
event but are observable in the spectroscopic survey and (2) the spectroscopic redshift determination of a large sample of galax-
number of observable galaxies in the error box may be too large ies at faint apparent magnitude (mr ≈ 24), as required by our
anyway (>104 ) to efficiently complete a spectroscopic survey study. The future spectroscopic survey BigBOSS (Schlegel et al.
on the full sample: self-similarity allows us to get the clus- 2009) is expected to further improve such figures of merit; a new

6
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

spectrograph will be able to simultaneously get up to 4000 spec- 5. CONSTRAINTS ON THE DARK ENERGY
tra within a single pointing of a 7 deg2 FOV. Getting a few thou- EQUATION OF STATE
sand spectra of objects falling within the GW error box in the
redshift range of interest may be possible in a single observing In this section, we present the results of our simulations.
night. At a mr = 24 cutoff magnitude we generally have few We tried several setups of the experiment by using different
hundred to few thousands galaxies in the GW error box, but thresholds on the observable apparent magnitude of galaxies,
we go deeper (i.e., mr = 26, feasible with future surveys), the different prescriptions for the measurement errors, and different
number of spectra may increase drastically. For some of the cosmological priors. For each setup, we performed either 100
error boxes, we count up to 105 galaxies with mr < 26. How- or 20 realizations of the MBH binary population as observed by
ever, the requirement of a factor of 10 more spectra does not LISA, together with the follow-up spectroscopic survey of the
correspond to a significant improvement of the results. This is a galaxies in all the error boxes.
consequence of the self-similarity of the galaxy distribution: as
long as there are enough galaxies in the error box to recover the 5.1. Fiducial Case
clustering information, the results are basically independent on
We consider in this subsection 100 realizations which we refer
the assumed cutoff magnitude. A survey with a cutoff magnitude
to as our fiducial case. For this setup, we limit spectroscopic
of mr = 24 may indeed be a good compromise between relia-
identification of galaxies in the error box to an apparent
bility of the results and time optimization in terms of follow-up
magnitude of mr  24, the errors in sky localization and in the
spectroscopy.
luminosity distance are estimated according to the inspiral part
The magnitude cutoff defines the number of neighbor ob-
of GW signal only, and the WL uncertainty is taken from Shapiro
servable galaxies. This is the only practical way to weight each
et al. (2010). The prior p0 (w) was assumed to be uniform
galaxy with a local density, ngal ≡ nmr (the subscript mr refers
U [−0.3:0.3] with an exponential decay at the boundaries. Such
to the adopted magnitude limit) along the lines discussed in
interval is consistent with current 2σ (95% confidence level)
Section 3.2. Once we have a spectroscopic galaxy sample, each
constraints on w (w = −0.12 ± 0.27; Komatsu et al. 2011),
galaxy in the error box comes with the prior probability to
obtained by cross-correlating seven year WMAP data with priors
be the host proportional to nmr , so the astrophysical prior in
coming from independent measurements of H0 and barionic
Equation (6) could be written as
acoustic oscillations (see Komatsu et al. 2011, and references
p(Ω, z) = nmr ,i δ(Ω − Ωi )δ(z − zi ), (11) therein for full details), under our same assumption for the
i
dark energy equation of state, ω = −1 − w, where w is a
constant. Such range is reduced by a factor of almost three
where the sum is over all observable galaxies in the error box (w = −0.02 ± 0.1) when SNe Ia data (Riess et al. 1998)
and Ω is the geodesic distance on the celestial sphere from are included. Here, we show that GW measurements offer
the center of the box. At redshifts z  1 the prior probability a competitive alternative to SNe Ia, placing an independent
p(Ω, z) becomes almost a continuous function (as the example constraint on the dark energy equation of state.
in Figure 4). We find that in almost all cases we improve the constraints
4.3. Approximations and Caveats on w, in other words, the posterior distribution is narrower
than the prior. Few events at low redshift usually play a major
Before discussing the results, we want to mention some role in the final result. One typical realization is plotted in
corrections we made to accommodate the limitations of our the top panel of Figure 5. We split the contribution to the
simulations. First, we interpreted one of the directions in the posterior distribution P (w) in redshift bands: z ∈ [0:1] (second
snapshot (along the side of the cylinder) as distance from the plot from the left), [1:2] (third plot), and [2:3] (fourth plot).
observer. This is a good approximation only if the error box Their relative contribution and the resulting posterior (black)
size is small. For large error boxes, a uniform distribution are given in the leftmost plot. In this example, the final posterior
in the comoving distances does not translate into a uniform probability is almost completely determined by few events
distribution in redshifts: there is an artificial slope with a bias at low redshift. The second realization, shown in the lower
toward low values of z. We have corrected for this slope. Second, panels of Figure 5, demonstrates how low-redshift contributions
the clumpiness evolves with redshift, which is not the case if could give inconclusive results. In this particular case, there
we use a single snapshot and interpret one of the directions as a are two maxima with preference given to the wrong one. The
redshift. To properly account for this, we should glue snapshots contribution from high-redshift events could change this ratio
together and perform an interpolation between them. However, as it is shown in this example. In many cases, the mergers above
we wanted to simplify the setup for this very first attempt. The redshift z = 1 can constrain w only to a 0.1–0.15 accuracy, but
main idea was to check whether the density contrast within the they almost always add up coherently, giving a maximum at the
error boxes is sufficient to further constrain the error on w. right value (w = 0). This usually helps in case the low-redshift
If the distribution of density within the error box is uniform events return a multimodal P (w), and is, in turn, the power of
then we do not gain any useful information. However, there our statistical method.
is a natural bias: for a given measurement of DL , the galaxy We characterize the results of each setup (100 or 20 realiza-
further away (larger z) constrains w better than a galaxy at lower tions) using the figures of merit shown in Figures 6 and 7. The
redshift. One can see it from the fact that deviation between the first one (Figure 6) is obtained by adding the posterior distribu-
curves in DL –z plane corresponding to the small deviation in tions P (w) of all the realizations. We fit the resulting curve with
w is larger for large z. This could be counterbalanced by the a Gaussian, characterizing the result using its mean w0 and stan-
decreasing density contrast at large redshift. Here, we corrected dard deviation σw . The second figure of merit (Figure 7) shows
the slope of the posterior Pj (w|s) by demanding that a uniform the result of Gaussian fits performed on each individual realiza-
distribution pj (θ, φ, z) returns a posterior on w equal to the tion (vertical index i): the mean w0 (i) is shown as a circle and the
prior, i.e., Pj (w|s) = p0 (w). standard deviation σw (i) is the error bar. The first figure of merit

7
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

0.08 0.02 0.01


0.08
0.008
0.06 0.015
0.06
0.006
0.04 0.01
0.04
0.004

0.02 0.02 0.005


0.002

0 0 0 0
-0.3 -0.15 0 0.15 0.3
z = [0:1] z = [1:2] z = [2:3]
0.05 0.05 0.02 0.01

0.04 0.04 0.008


0.015

0.03 0.03 0.006


0.01
0.02 0.02 0.004

0.005
0.01 0.01 0.002

0 0 0 0
-0.3 -0.15 0 0.15 0.3 -0.3 -0.15 0 0.15 0.3 -0.3 -0.15 0 0.15 0.3 -0.3 -0.15 0 0.15 0.3
w w w w
Figure 5. Posterior distribution for w for two particular realizations (top and bottom row). In each row, the left plot shows the full posterior from all GW events (black
curve) as well as contributions from different redshift bands. The three right plots show the individual contribution for the three redshift ranges, as labeled in the panels.
(A color version of this figure is available in the online journal.)

(a) fiducial case (b) without electromagnetic counterpart (c) with offset gaussian prior
10 9 12
9 Data 8 Data Data
Gaussian fit : Gaussian fit : 10 Gaussian fit :
8 w0=0.0008 7 w0=0.0007 w0=0.0069
7 σ=0.0359 6 σ=0.0397 8 σ=0.0336
6
5
5 6
4
4
3 3 4
2 2
2
1 1
0 0 0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
w w w
(d) apparent magnitude <26 (e) with merger (f) improved WL + merger
12 12 18
Data Data 16 Data
10 Gaussian fit : 10 Gaussian fit : Gaussian fit :
w0=0.0058 w0=-0.0067 14 w0=-0.0025
8 σ=0.0329 8 σ=0.0357 12 σ=0.0213
10
6 6
8
4 4 6
4
2 2
2
0 0 0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 -0.3 -0.2 -0.1 0 0.1 0.2 0.3
w w w

Figure 6. Collective figures of merit of our experiment. In each panel, corresponding to a different setup of our experiment as labeled in figure, the red solid curve
corresponds to the data, i.e., the sum of the posterior distributions of w over all realizations. The blue dashed curve is a Gaussian fit with parameter given in the legend
of each plot.
(A color version of this figure is available in the online journal.)

gives collective information, showing how well, on average, an corresponding to a factor of four improvement in the estimation
individual realization can be approximated by a Gaussian fit, of w with respect to our standard 2σ [−0.3:0.3] prior. However,
while the second figure of merit shows the dispersion of the the distribution has clearly some outliers, recognizable as non-
posterior distribution across the individual realizations. Gaussian tails in Figure 6 and pinned down in Figure 7. For the
The fiducial case, featuring 100 realizations, is shown in panel fiducial case, 84% of the realizations have a mean value close
(a) of both Figures 6 and 7. The parameters of the global to the true one, i.e., |w0 (i) − wtrue | < 0.1 with an appreciable
fitting Gaussian mean are w0 = 0.0008 and σw = 0.036, reduction of the prior range, i.e., σw (i) < 0.15 (i = 1, . . . , 100

8
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

(a) fiducial case (d) apparent magnitude <26


100 20

80
15
60

index

index
10
40
5
20

0 0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 -0.3 -0.2 -0.1 0 0.1 0.2 0.3
(e) with merger (f) improved WL + merger
20 20

15 15
index

index
10 10

5 5

0 0
-0.2 0 0.2 -0.3 -0.2 -0.1 0 0.1 0.2 0.3
w, δw w, δw
Figure 7. Mean values and standard deviations resulting from the Gaussian fit of the posterior P (w). The setup of each panel corresponds to the one adopted in the
panel of Figure 6 labeled by the same letter).

is the realization index). Moreover, most of the outliers can be mr = 26 is given in panel (d) of Figures 6 and 7. The results are
corrected as we will explain in Section 5.6. comparable to the fiducial case. They show a small improvement
in sigma and slightly larger bias for the combined distribution.
5.2. Removing “Electromagnetic Counterparts” We also notice that four out of five “bad” cases remain bad.
Our goal is to demonstrate that we are able to constrain We should say few words about the number of galaxies used
the dark energy equation of state without directly observing here. As mentioned above, the typical number of galaxies for
electromagnetic counterparts. However, for some of the low- the fiducial case (mr = 24) is less than few thousand for events
redshift events, the error box is so small that only one or two at z < 1 and less than few tens of thousands for the high-redshift
galaxies fall within it. Having one or two galaxies in the error event. For the improved observational limit (mr = 26), these
box essentially implies an electromagnetic identification of the numbers are 2–10 times larger. The fact that our results are not
host, so we decided to re-analyze the fiducial case removing sensitive to the depth of the survey reflects the self-similarity of
all such fortunate events (usually 0–2 in each realization). The the spatial distribution of galaxies in different mass ranges.
fiducial case without clearly identifiable hosts is presented in 5.5. Improving the Sky Localization and the Luminosity
the panel (b) of Figure 6. Clearly, our results remain almost Distance Estimation
unchanged, the posterior distribution is slightly wider (larger
sigma) and non-Gaussianity is more pronounced. In our fiducial setup, the assumed source sky localization and
luminosity distance error are rather conservative. In this subsec-
5.3. Choice of the Prior for w tion, we consider the effect of improving such measurements.
So far, we considered only the inspiral part of the GW signal;
Here and in the next subsections, we make use of 20 selected the inclusion of merger and ringdown will improve the localiza-
realizations, which we found to be sufficient to depict the tion of the source by at least a factor of two (McWilliams et al.
relevant trends of the analysis. We took 15 “good” (mean values 2010), due to the large gain in S/N. We artificially reduced the
close to the true and small rms errors) and 5 “bad” cases from sky localization error coming from the inspiral by a factor of
the fiducial setup. two (factor of four in the area), assuming that this will be the
In this subsection, we study the effect of the prior p0 (w) on case if we take the full GW signal. We re-analyzed the same
the posterior distribution. We considered an extreme case: a 20 realizations with this new error on the sky. Because the size
Gaussian N (w0 = −0.2, σ = 0.3). As shown in panel (c) of of the error box is smaller, the number of potential counter-
Figure 6, the global posterior distribution is still centered at the parts is reduced by a factor of ∼4 compared with the fiducial
true value w = 0. This demonstrates that the final conclusion case. The results are presented in panel (e) of Figure 6. We see
is basically unaffected by the choice of the prior (as long as the that the main effect of a better GW source localization is to re-
prior covers the true value) and GW observations, in principle, duce the number of outliers and to remove the non-Gaussian tails
could be used as an independent mean of estimating w. in the combined probability. As it is clear from panel (e) of
5.4. Using Deeper Surveys Figure 7, the main gain comes from improvement of the “bad”
cases.
Here, we study the dependence of our results on the depth We now consider another estimation of the mean WL con-
of the follow-up spectroscopic survey: i.e., on the observability tribution to the luminosity distance error, given in Wang et al.
threshold. We considered the same 20 realizations as in the (2002; green square-line curve on Figure 2). We take this in
previous section, but now with different limits on the apparent combination with improved source localization on the sky com-
magnitude of observable galaxies: mr = 24, 25, 26. The case ing from taking into account the merger (as discussed above).

9
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
-0.4 -0.2 0 0.2 0.4 -0.4 -0.2 0 0.2 0.4

0.03 0.03

0.025

0.02 0.02

0.015

0.01 0.01

0.005

0 0
-0.4 -0.2 0 0.2 0.4 -0.4 -0.2 0 0.2 0.4
w w
Figure 8. Each row of panels shows our self-similarity check for a selected realization. In each row, the solid curve on the left panel corresponds the final posterior

k (w).
P (w). The solid curves on the right panel are the posteriors after removing one event, P
(A color version of this figure is available in the online journal.)

We consider the same 20 realizations. Results are shown in panel removing one event at low redshift changes the final probability
(f) of both Figures 6 and 7. The improvement with respect to all completely; the solid (red) line in the right panel is the new
the other cases is obvious. Because the marginalized likelihood posterior distribution, consistent with the true value w = 0.
πj coming with each galaxy is narrower due to the smaller error However, there are still few cases where the self-consistency
in the luminosity distance, the final posterior on Pj (w) is also test is not conclusive, and one of them is shown in the lower
narrower. The standard deviation σw is improved by more than panels of Figure 8. In this case, removing one “bad” nearby
40% as compared with the fiducial case. The non-Gaussian tails event produces the red curve centered at w = 0, but removing
have almost completely disappeared, due to the removal of the another (“good”) event results in the green curve, which are
outliers (further improvement of the “bad” cases, the remain- mutually not consistent at all. Since in real life we will not
ing bad case will be treated in the Section 5.6; see also the top know which event is “good” and which one is “bad,” we will
panel of Figure 8). With this model of the mean WL contribu- not be able to make a clear definite statement, and our answer
tion and assuming the full GW signal, the estimation of w is will be bi-modal with a probability attached to each mode.
improved by a factor of ∼8 as compared to the initial uniform
prior. 5.7. Comparison with the Optimal Case: Detection of
Electromagnetic Counterparts
5.6. Consistency Check
For comparison, we have also considered the best possible
As we mentioned above, some nearby GW event could seri- case, in which the redshifts of the GW source hosts are de-
ously bias the final posterior. We also mentioned that the odds termined unambiguously through the identification of a dis-
for the host to be in a low-density region of the universe are not tinctive electromagnetic counterpart. In this case, the redshift
small. The posterior probability P (w) reflects the distribution of each GW event is known exactly (within negligible mea-
of the mass defined by the astrophysical prior pj (θ, φ, z). A surement errors). Therefore, the error on w comes only from
nearby GW event hosted in the low-density environment could the error on luminosity distance (GW error measurement plus
seriously damage the final result. An example is given in the top WL). Considering 20 realizations with a configuration equiva-
left panel of Figure 8. In order to eliminate or at least test such lent to the fiducial case (Section 5.1), the global posterior dis-
unfortunate cases, we performed a self-consistency test on our tribution is a Gaussian centered at w0 = 0 with σw = 0.021
results. Basically, we remove one GW event from the analysis (for comparison, see panel (a) of Figure 6). With a con-
and see if the resulting posterior P
k (w) distributions are con- figuration equivalent to our improved case, i.e., better WL
sistent. We defined the posterior of all the events minus one as (Section 5.5), we obtain σw = 0.012 (for comparison, see to
 panel (f) of Figure 6). In both cases, the difference between our
p0 (w) j
=k Pj (s|w)
P
k (w) =   . (12) statistical method and the best possible case (all electromagnetic
p0 (w) j
=k Pj (s|w)dw counterparts detected) is only about a factor two.

k (w) gives similar results for all k, then we can be confident
If P 6. SUMMARY
that the result is not biased by one particular unfortunate event,
and this increases our trust in the final posterior distribution. In this paper, we presented a statistical method for con-
If, conversely, all P
k (w) but one are consistent, then we say straining cosmological parameters using LISA observations of
that this one event is not in line with the remaining events and spinning MBH binaries and redshift surveys of galaxies. Our ap-
should be abandoned. In the top panels of Figure 8, we see that proach does not require any direct electromagnetic counterpart;

10
The Astrophysical Journal, 732:82 (11pp), 2011 May 10 Petiteau, Babak, & Sesana

instead, the consistency between a few dozen GW events im- parameters, testing LISA capabilities of setting constraints on
poses constraints on the redshift–luminosity distance relation- a multi-parameter model.
ship. This, in turn, allows us to estimate cosmological param-
eters. This method strongly relies on the non-uniformity (i.e., The work of A.P. and S.B. was supported in parts by
clustering) of the galaxy distribution within the uncertainty error DFG grant SFB/TR 7 Gravitational Wave Astronomy and by
box set by LISA observations, WL and priors on the cosmolog- DLR (Deutsches Zentrum fur Luft-und Raumfahrt). The Monte
ical parameters. Carlo simulations were performed on the Morgane cluster at
For this first exploratory study, we fixed all the cosmological AEI-Golm and on the Atlas cluster at AEI-Hannover. The
parameters but one, w, describing the effective equation of state authors thank Jonathan Gair and Toshifumi Futamase for useful
for the dark energy. We used the Millennium simulation to model discussions.
the universe at different redshifts. We used a particular (VB)
hierarchical MBH formation model to mimic the MBH binary REFERENCES
population observed by LISA. Using this setup, we considered
between 20 and 100 realizations of the observed LISA binary Arun, K. G., Iyer, B. R., Sathyaprakash, B. S., Sinha, S., & Van Den Broeck, C.
population. We tried two different models for estimating the 2007, Phys. Rev., D76, 104016
Arun, K. G., Mishra, C. K., Van Den Broeck, C., Iyer, B. R., Sathyaprakash,
error in luminosity distance due to WL, we also looked at the B. S., & Sinha, S. 2009a, Class. Quantum Gravity, 26, 094021
effect of including merger and ringdown via improvement of Arun, K. G., et al. 2009b, Class. Quantum Grav., 26, 094027
the sky localization. We checked the robustness of our final result Babak, S., et al. 2010, Class. Quantum Gravity, 27, 084009
against different depth of future spectroscopic galaxy surveys. Begelman, M. C., Volonteri, M., & Rees, M. J. 2006, MNRAS, 370, 289
Bertone, S., De Lucia, G., & Thomas, P. A. 2007, MNRAS, 379, 1143
Our fiducial case, based on conservative assumptions, shows Bielby, R., et al. 2010, arXiv:1005.3028
that we are able to constrain w to a 8% level (2σ ), i.e., we Bower, R. G., Benson, A. J., Malbon, R., Helly, J. C., Frenk, C. S., Baugh,
improve its estimate by a factor of ∼4 as compared to the C. M., Cole, S., & Lacey, C. G. 2006, MNRAS, 370, 645
current 95% confidence interval obtained by cross-correlating Cornish, N. J., & Porter, E. K. 2007, Class. Quantum Grav., 24, 5729
the seven year WMAP data analysis with priors coming from Danzmann, K., & the LISA Study Team 1997, Class. Quantum Gravity, 14,
1399
H0 measurements and barionic acoustic oscillations (Komatsu Davis, M., et al. 2003, Proc. SPIE, 4834, 161
et al. 2011). Such new measurement would be at the same De Lucia, G., & Blaizot, J. 2007, MNRAS, 375, 2
level (25% better on average) than current constraints based on Faber, S. M., et al. 2003, Proc. SPIE, 4841, 1657
seven year WMAP data plus SNe Ia observations. The optimistic Gair, J. R., Sesana, A., Berti, E., & Volonteri, M. 2010, arXiv:1009.6172
Gültekin, K., et al. 2009, ApJ, 698, 198
case (smaller WL disturbance and full GW waveform) allows Haehnelt, M. G., & Rees, M. J. 1993, MNRAS, 263, 168
us a further improvement by another factor of two, providing Hilbert, S., Gair, J. R., & King, L. J. 2011, MNRAS, 412, 1023
a factor of ∼2.5 tighter constraint than current estimates Holz, D. E., & Hughes, S. A. 2005, ApJ, 629, 15
including SNe data. Our results are most sensitive to the WL Komatsu, E., et al. 2011, ApJS, 192, 18
error (witnessing once more how critical is the issue of WL Lang, R. N., & Hughes, S. A. 2006, Phys. Rev. D, 74, 122001
Lang, R. N., & Hughes, S. A. 2009, Class. Quantum Grav., 26, 094035
mitigation for cosmological parameter estimation through GW Lang, R. N., Hughes, S. A., & Cornish, N. J. 2011, arXiv:1101.3591
observations) and are almost independent on the depth of the Le Fèvre, O., et al. 2003, Proc. SPIE, 4841, 1670
redshift survey (provided we have a reasonable number of Le Fèvre, O., et al. 2005, A&A, 439, 845
redshift measurements per error box). Lilly, S. J., et al. 2009, ApJS, 184, 218
MacLeod, C. L., & Hogan, C. J. 2008, Phys. Rev. D, 77, 043512
In the majority of the realizations the most information Madau, P., & Rees, M. J. 2001, ApJ, 551, L27
comes from few events at low redshift, and high-redshift events Magorrian, J., et al. 1998, AJ, 115, 2285
do help in case of multimodal structures in the posterior Malbon, R. K., Baugh, C. M., Frenk, C. S., & Lacey, C. G. 2007, MNRAS, 382,
distribution. We suggested a self-consistency check based on 1394
the similarity of the posterior distribution from each GW event. McWilliams, S. T., Thorpe, J. I., Baker, J. G., & Kelly, B. J. 2010, Phys. Rev. D,
81, 064014
This increases our confidence in the final result and allows a Oke, J. B., & Sandage, A. 1968, ApJ, 154, 21
reduction of the risk of incurring unfortunate outlier realizations Petiteau, A., Shang, Y., Babak, S., & Feroz, F. 2010, Phys. Rev. D, 81, 104016
for which we cannot place useful constraints on w. We also Plowman, J. E., Hellings, R. W., & Tsuruta, S. 2010, arXiv:1009.0765
compared our statistical method to the optimal situation in which Riess, A. G., et al. 1998, AJ, 116, 1009
Salvaterra, R., Haardt, F., & Volonteri, M. 2007, MNRAS, 374, 761
electromagnetic counterparts to the GW sources are identified, Schlegel, D. J., et al. 2009, arXiv:0904.0468
finding an improvement of a factor of two in the latter case. In Schnittman, J. D. 2010, arXiv:1010.3250
absence of distinctive electromagnetic counterparts, statistical Schutz, B. F. 1986, Nature, 323, 310
methods like the one presented here can still efficiently constrain Sesana, A., Gair, J. R., Berti, E., & Volonteri, M. 2010, Phys. Rev. D, 83, 4036
cosmological parameters. Shapiro, C., Bacon, D. J., Hendry, M., & Hoyle, B. 2010, MNRAS, 404, 858
Springel, V., et al. 2005, Nature, 435, 629
Although the main result of the present paper is encouraging, Trias, M., & Sintes, A. M. 2008, Phys. Rev., D77, 024030
it was obtained assuming a fixed cosmological model with one Vallisneri, M. 2008, Phys. Rev. D, 77, 042001
free parameter only: the w parameter describing the dark energy Van Den Broeck, C., Trias, M., Sathyaprakash, B. S., & Sintes, A. M. 2010,
equation of state. Even though we will likely have a good Phys. Rev., D81, 124031
Volonteri, M., & Begelman, M. C. 2010, MNRAS, 409, 1022
knowledge of most of the other cosmological parameters by Volonteri, M., Haardt, F., & Madau, P. 2003, ApJ, 582, 559
the time LISA will fly, it is worth considering models with more Wang, Y., Holz, D. E., & Munshi, D. 2002, ApJ, 572, L15
degrees of freedom. In following studies, we intend to consider White, S. D. M., & Rees, M. J. 1978, MNRAS, 183, 341
a more realistic situation by releasing other cosmological Zombeck, M. V. 1990, Science, 249, 1314

11

You might also like